You are on page 1of 15

Polymer

From Wikipedia, the free encyclopedia

(Redirected from Polymers)


Jump to: navigation, search

Appearance of real linear polymer chains as recorded using an atomic force microscope
on surface under liquid medium. Chain contour length for this polymer is ~204 nm;
thickness is ~0.4 nm.[1]

A polymer (from Greek πολύ-ς /po΄li-s/ much, many and μέρος /΄meros/ part) is a large
molecule (macromolecule) composed of repeating structural units typically connected by
covalent chemical bonds. While polymer in popular usage suggests plastic, the term
actually refers to a large class of natural and synthetic materials with a variety of
properties.

Due to the extraordinary range of properties accessible in polymeric materials [2], they
have come to play an essential and ubiquitous role in everyday life[3] - from plastics and
elastomers on the one hand to natural biopolymers such as DNA and proteins that are
essential for life on the other. A simple example is polyethylene, whose repeating unit is
based on ethylene (IUPAC name ethene) monomer. Most commonly, as in this example,
the continuously linked backbone of a polymer consists mainly of carbon atoms.
However, other structures do exist; for example, elements such as silicon form familiar
materials such as silicones, examples being silly putty and waterproof plumbing sealant.
The backbone of DNA is in fact based on a phosphodiester bond, and repeating units of
polysaccharides (e.g. cellulose) are joined together by glycosidic bonds via oxygen
atoms.

Natural polymeric materials such as shellac, amber, and natural rubber have been in use
for centuries. Biopolymers such as proteins and nucleic acids play crucial roles in
biological processes. A variety of other natural polymers exist, such as cellulose, which is
the main constituent of wood and paper.

The list of synthetic polymers includes synthetic rubber, Bakelite, neoprene, nylon, PVC,
polystyrene, polyacrylonitrile, PVB, silicone, and many more.
Polymers are studied in the fields of polymer chemistry, polymer physics, and polymer
science.

Contents
[hide]

• 1 Etymology
• 2 Historical development
• 3 Polymer synthesis
o 3.1 Laboratory synthesis
o 3.2 Biological synthesis
o 3.3 Modification of natural polymers
• 4 Polymer properties
o 4.1 Monomers / Repeat Units
o 4.2 Microstructure
 4.2.1 Polymer Architecture
 4.2.2 Chain length
 4.2.3 Monomer arrangement in copolymers
 4.2.4 Tacticity
o 4.3 Polymer Morphology
 4.3.1 Crystallinity
 4.3.2 Chain conformation
o 4.4 Mechanical Properties
 4.4.1 Tensile strength
 4.4.2 Young's modulus of elasticity
 4.4.3 Transport properties
o 4.5 Phase behavior
 4.5.1 Melting point
 4.5.2 Boiling point
 4.5.3 Glass transition temperature
 4.5.4 Mixing behavior
 4.5.5 Inclusion of plasticizers
o 4.6 Chemical properties
• 5 Standardized polymer nomenclature
• 6 Polymer characterization
• 7 Polymer degradation
o 7.1 Product failure
• 8 References
• 9 Bibliography
• 10 See also

• 11 External links

[edit] Etymology
The word polymer is derived from the Greek words πολυ (poly), meaning "many"; and
μέρος (meros), meaning "part". The term was coined in 1833 by Jöns Jakob Berzelius,
although his definition of a polymer was quite different from the modern definition. (see
Jöns Jakob Berzelius#New chemical terms)

[edit] Historical development


Starting in 1811, Henri Braconnot did pioneering work in derivative cellulose
compounds, perhaps the earliest important work in polymer science. The development of
vulcanization later in the nineteenth century improved the durability of the natural
polymer rubber, signifying the first popularized semi-synthetic polymer. In 1907, Leo
Baekeland created the first completely synthetic polymer, Bakelite, by reacting phenol
and formaldehyde at precisely controlled temperature and pressure. Bakelite was then
publicly introduced in 1909.

Despite significant advances in synthesis and characterization of polymers, a correct


understanding of polymer molecular structure did not emerge until the 1920s. Before
then, scientists believed that polymers were clusters of small molecules (called colloids),
without definite molecular weights, held together by an unknown force, a concept known
as association theory. In 1922, Hermann Staudinger proposed that polymers consisted of
long chains of atoms held together by covalent bonds, an idea which did not gain wide
acceptance for over a decade and for which Staudinger was ultimately awarded the Nobel
Prize. Work by Wallace Carothers in the 1920s also demonstrated that polymers could be
synthesized rationally from their constituent monomers. An important contribution to
synthetic polymer science was made by the Italian chemist Giulio Natta and the German
chemist Karl Ziegler, who won the Nobel Prize in Chemistry in 1963 for the development
of the Ziegler-Natta catalyst. Further recognition of the importance of polymers came
with the award of the Nobel Prize in Chemistry in 1974 to Paul Flory, whose extensive
work on polymers included the kinetics of step-growth polymerization and of addition
polymerization, chain transfer, excluded volume, the Flory-Huggins solution theory, and
the Flory convention.

Synthetic polymer materials such as nylon, polyethylene, Teflon, and silicone have
formed the basis for a burgeoning polymer industry. These years have also shown
significant developments in rational polymer synthesis. Most commercially important
polymers today are entirely synthetic and produced in high volume on appropriately
scaled organic synthetic techniques. Synthetic polymers today find application in nearly
every industry and area of life. Polymers are widely used as adhesives and lubricants, as
well as structural components for products ranging from children's toys to aircraft. They
have been employed in a variety of biomedical applications ranging from implantable
devices to controlled drug delivery. Polymers such as poly(methyl methacrylate) find
application as photoresist materials used in semiconductor manufacturing and low-k
dielectrics for use in high-performance microprocessors. Recently, polymers have also
been employed as flexible substrates in the development of organic light-emitting diodes
for electronic displays.
[edit] Polymer synthesis
Main article: Polymerization

The repeating unit of the polymer polypropylene

Polymerization is the process of combining many small molecules known as monomers


into a covalently bonded chain. During the polymerization process, some chemical
groups may be lost from each monomer. This is the case, for example, in the
polymerization of PET polyester. The monomers are terephthalic acid (HOOC-C6H4-
COOH) and ethylene glycol (HO-CH2-CH2-OH) but the repeating unit is -OC-C6H4-
COO-CH2-CH2-O-, which corresponds to the combination of the two monomers with the
loss of two water molecules. The distinct piece of each monomer that is incorporated into
the polymer is known as a repeat unit or monomer residue.

[edit] Laboratory synthesis

Laboratory synthetic methods are generally divided into two categories, step-growth
polymerization and chain-growth polymerization[4]. The essential difference between the
two is that in chain growth polymerization, monomers are added to the chain one at a
time only[5], whereas in step-growth polymerization chains of monomers may combine
with one another directly[6]. However, some newer methods such as plasma
polymerization do not fit neatly into either category. Synthetic polymerization reactions
may be carried out with or without a catalyst. Efforts towards rational synthesis of
biopolymers via laboratory synthetic methods, especially artificial synthesis of proteins,
is an area of intense research.

[edit] Biological synthesis

Main article: Biopolymer

There are three main classes of biopolymers: polysaccharides, polypeptides, and


polynucleotides. In living cells, they may be synthesized by enzyme-mediated processes,
such as the formation of DNA catalyzed by DNA polymerase. The synthesis of proteins
involves multiple enzyme-mediated processes to transcribe genetic information from the
DNA and subsequently translate that information to synthesize the specified protein from
amino acids. The protein may be modified further following translation in order to
provide appropriate structure and functioning.

[edit] Modification of natural polymers

Many commercially important polymers are synthesized by chemical modification of


naturally occurring polymers. Prominent examples include the reaction of nitric acid and
cellulose to form nitrocellulose and the formation of vulcanized rubber by heating natural
rubber in the presence of sulphur.

[edit] Polymer properties


Polymer properties are broadly divided into several classes based on the scale at which
the property is defined as well as upon its physical basis. The most basic property of a
polymer is the identity of its constituent monomers. A second set of properties, known as
microstructure, essentially describe the arrangement of these monomers within the
polymer at the scale of a single chain. These basic structural properties play a major role
in determining bulk physical properties of the polymer, which describe how the polymer
behaves as a continuous macroscopic material. Chemical properties, at the nano-scale,
describe how the chains interact through various physical forces. At the macro-scale, they
describe how the bulk polymer interacts with other chemicals and solvents.

[edit] Monomers / Repeat Units

The identity of the monomer residues (repeat units) comprising a polymer is its first and
most important attribute. Polymer nomenclature is generally based upon the type of
monomer residues comprising the polymer. Polymers that contain only a single type of
repeat unit are known as homopolymers, while polymers containing a mixture of repeat
units are known as copolymers. Poly(styrene), for example, is composed only of styrene
monomer residues, and is therefore classified as a homopolymer. Ethylene-vinyl acetate,
on the other hand, contains more than one variety of repeat unit and is thus a copolymer.
Some biological polymers are composed of a variety of different but structurally related
monomer residues; for example, polynucleotides such as DNA are composed of a variety
of nucleotide subunits.

A polymer molecule containing ionizable subunits is known as a polyelectrolyte or


ionomer.

[edit] Microstructure

The microstructure of a polymer (sometimes called configuration) relates to the physical


arrangement of monomer residues along the backbone of the chain[7]. These are the
elements of polymer structure that require the breaking of a covalent bond in order to
change. Structure has a strong influence on the other properties of a polymer. For
example, two samples of natural rubber may exhibit different durability, even though
their molecules comprise the same monomers.

[edit] Polymer Architecture

Branch point in a polymer


An important microstructural feature determining polymer properties is the polymer
architecture.[8] The simplest polymer architecture is a linear chain: a single backbone
with not branches. A related unbranching architecture is a ring polymer. A branched
polymer molecule is composed of a main chain with one or more substituent side chains
or branches. Special types of branched polymers include star polymers, comb polymers,
brush polymers, ladders, and dendrimers[8].

Branching of polymer chains affects the ability of chains to slide past one another by
altering intermolecular forces, in turn affecting bulk physical polymer properties. Long
chain branches may increase polymer strength, toughness, and the glass transition
temperature due to an increase in the number of entanglements per chain. The effect of
such long-chain branches on the size of the polymer in solution is characterized by the
branching index. Random length and atactic short chains, on the other hand, may reduce
polymer strength due to disruption of organization and may likewise reduce the
crystallinity of the polymer.

A good example of this effect is related to the range of physical attributes of


polyethylene. High-density polyethylene (HDPE) has a very low degree of branching, is
quite stiff, and is used in applications such as milk jugs. Low-density polyethylene
(LDPE), on the other hand, has significant numbers of both long and short branches, is
quite flexible, and is used in applications such as plastic films.

Figure 1. Dendrimer and dendron

Dendrimers are a special case of polymer where every monomer unit is branched. This
tends to reduce intermolecular chain entanglement and crystallization. Alternatively,
dendritic polymers are not perfectly branched but share similar properties to dendrimers
due to their high degree of branching.

The architecture of the polymer is often physically determined by the functionality of the
monomers from which it is formed[9]. This property of a monomer is defined as the
number of reaction sites at which may form chemical covalent bonds. The basic
functionality required for forming even a linear chain is two bonding sites. Higher
functionality yields branched or even crosslinked or networked polymer chains.

An effect related to branching is chemical crosslinking - the formation of covalent bonds


between chains. Crosslinking tends to increase Tg and increase strength and toughness.
Among other applications, this process is used to strengthen rubbers in a process known
as vulcanization, which is based on crosslinking by sulphur. Car tires, for example, are
highly crosslinked in order to reduce the leaking of air out of the tire and to toughen their
durability. Eraser rubber, on the other hand, is not crosslinked to allow flaking of the
rubber and prevent damage to the paper.

A cross-link suggests a branch point from which four or more distinct chains emanate. A
polymer molecule with a high degree of crosslinking is referred to as a polymer network.
[10]
Sufficiently high crosslink concentrations may lead to the formation of an infinite
network, also known as a gel, in which networks of chains are of unlimited extent —
essentially all chains have linked into one molecule.[11]

[edit] Chain length

The physical properties of a polymer are strongly dependent on the size or length of the
polymer chain. [12]. For example, as chain length is increased, melting and boiling
temperatures increase quickly[12]. Impact resistance also tends to increase with chain
length, as does the viscosity, or resistance to flow, of the polymer in its melt state[13].
Chain length is related to melt viscosity roughly as 1:103.2, so that a tenfold increase in
polymer chain length results in a viscosity increase of over 1000 times[citation needed].
Increasing chain length furthermore tends to decrease chain mobility, increase strength
and toughness, and increase the glass transition temperature (Tg)[citation needed]. This is a
result of the increase in chain interactions such as Van der Waals attractions and
entanglements that come with increased chain length[citation needed]. These interactions tend to
fix the individual chains more strongly in position and resist deformations and matrix
breakup, both at higher stresses and higher temperatures[citation needed].

A common means of expressing the length of a chain is the degree of polymerization,


which quanitifies the number of monomers incorporated into the chain[14][15]. As with
other molecules, a polymer's size may also be expressed in terms of molecular weight.
Since synthetic polymerization techiniques typically yields a polymer product including a
range of molecular weights, the weight is often expressed statistically to describe the
distribution of chain lengths present in the same. Common examples are the number
average molecular weight and weight average molecular weight[16][17]. The ratio of these
two values is the polydispersity index, commonly used to express the "width" of the
molecular weight distribution[18]. A final measurement is contour length, which can be
understood as the length of the chain backbone in its fully extended state[19].

The flexibility of an unbranched chain polymer is characterized by its persistence length.

[edit] Monomer arrangement in copolymers

Main article: copolymer

Monomers within a copolymer may be organized along the backbone in a variety of


ways.
• Alternating copolymers possess regularly alternating monomer residues[20] (2).
• Periodic copolymers have monomer residue types arranged in a repeating
sequence. .
• Statistical copolymers have monomer residues arranged according to a known
statistical rule. A statistical copolymer in which the probability of finding a
particular type of monomer residue at an particular point in the chain is
independent of the types of surrounding monomer residue may be referred to as a
truly random copolymer[21][22] (3).
• Block copolymers have two or more homopolymer subunits linked by covalent
bonds[20] (4). Polymers with two or three blocks of two distinct chemical species
(e.g., A and B) are called diblock copolymers and triblock copolymers,
respectively. Polymers with three blocks, each of a different chemical species
(e.g., A, B, and C) are termed triblock terpolymers.
• Graft or grafted copolymers contain side chains that have a different
composition or configuration than the main chain.(5)

[edit] Tacticity

Main article: Tacticity

Tacticity describes the relative stereochemistry of chiral centers in neighboring structural


units within a macromolecule. There are three types: isotactic (all substituents on the
same side), atactic (random placement of substituents), and syndiotactic (alternating
placement of substituents).

[edit] Polymer Morphology

Polymer morphology generally describes the arrangement of chains in space and


microscopic ordering of many polymer chains.

[edit] Crystallinity

When applied to polymers, the term crystalline has a somewhat ambiguous usage. In
some cases, the term crystalline finds identical usage to that used in conventional
crystallography. For example, the structure of a crystalline protein or polynucleotide,
such as a sample prepared for x-ray crystallography, may be defined in terms of a
conventional unit cell composed of one or more polymer molecules with cell dimensions
of hundreds of angstroms or more.

A synthetic polymer may be lightly described as crystalline if it contains regions of three-


dimensional ordering on atomic (rather than macromolecular) length scales, usually
arising from intramolecular folding and/or stacking of adjacent chains. Synthetic
polymers may consist of both crystalline and amorphous regions; the degree of
crystallinity may be expressed in terms of a weight fraction or volume fraction of
crystalline material. Few synthetic polymers are entirely crystalline.[23]
The crystallinity of polymers is characterized by their degree of crystallinity, ranging
from zero for a completely noncrystalline polymer to one for a theoretical completely
crystalline polymer. Increasing degree of crystallinity tends to make a polymer more
rigid. It can also lead to greater brittleness. Polymers with a degree of crystallinity
approaching zero or one will tend to be transparent, while polymers with intermediate
degrees of crystallinity will tend to be opaque due to light scattering by crystalline or
glassy regions. Thus for many polymers, reduced crystallinity may also be associated
with increased transparency.

[edit] Chain conformation

The space occupied by a polymer molecule is generally expressed in terms of radius of


gyration, which is an average distance from the center of mass of the chain to the chain
itself. Alternatively, it may be expressed in terms of pervaded volume, which is the
volume of solution spanned by the polymer chain and scales with the cube of the radius
of gyration.[24]

[edit] Mechanical Properties

The bulk properties of a polymer are those most often of end-use interest. These are the
properties that dictate how the polymer actually behaves on a macroscopic scale.

[edit] Tensile strength

The tensile strength of a material quantifies how much stress the material will endure
before failing.[25][26] This is very important in applications that rely upon a polymer's
physical strength or durability. For example, a rubber band with a higher tensile strength
will hold a greater weight before snapping. In general tensile strength increases with
polymer chain length and crosslinking of polymer chains.

[edit] Young's modulus of elasticity

Young's Modulus quantifies the elasticity of the polymer. It is defined, for small strains,
as the ratio of rate of change of stress to strain. Like tensile strength, this is highly
relevant in polymer applications involving the physical properties of polymers, such as
rubber bands. The modulus is strongly dependent on temperature.

[edit] Transport properties

Transport properties such as diffusivity relate to how rapidly molecules move through the
polymer matrix. These are very important in many applications of polymers for films and
membranes.

[edit] Phase behavior

[edit] Melting point


The term melting point, when applied to polymers, suggests not a solid-liquid phase
transition but a transition from a crystalline or semi-crystalline phase to a solid
amorphous phase. Though abbreviated as simply Tm, the property in question is more
properly called the crystalline melting temperature. Among synthetic polymers,
crystalline melting is only discussed with regards to thermoplastics, as thermosetting
polymers will decompose at high temperatures rather than melt.

[edit] Boiling point

The boiling point of a polymeric material is strongly dependent on chain length. High
polymers with a large degree of polymerization do not exhibit a boiling point because
they decompose before reaching theoretical boiling temperatures. For shorter oligomers,
a boiling transition may be observed and will generally increase rapidly as chain length is
increased.

[edit] Glass transition temperature

A parameter of particular interest in synthetic polymer manufacturing is the glass


transition temperature (Tg), which describes the temperature at which amorphous
polymers undergo a second-order phase transition from a rubbery, viscous amorphous
solid, or from a crystalline solid (depending on the degree of crystallization) to a brittle,
glassy amorphous solid. The glass transition temperature may be engineered by altering
the degree of branching or crosslinking in the polymer or by the addition of plasticizer.[27]

[edit] Mixing behavior

In general, polymeric mixtures are far less miscible than mixtures of small molecule
materials. This effect results from the fact that the driving force for mixing is usually
entropy, not interaction energy. In other words, miscible materials usually form a solution
not because their interaction with each other is more favorable than their self-interaction,
but because of an increase in entropy and hence free energy associated with increasing
the amount of volume available to each component. This increase in entropy scales with
the number of particles (or moles) being mixed. Since polymeric molecules are much
larger and hence generally have much higher specific volumes than small molecules, the
number of molecules involved in a polymeric mixture are far less than the number in a
small molecule mixture of equal volume. The energetics of mixing, on the other hand, are
comparable on a per volume basis for polymeric and small molecule mixtures. This tends
to increase the free energy of mixing for polymer solutions and thus make solvation less
favorable. Thus, concentrated solutions of polymers are far rarer than those of small
molecules.

In dilute solution, the properties of the polymer are characterized by the interaction
between the solvent and the polymer. In a good solvent, the polymer appears swollen and
occupies a large volume. In this scenario, intermolecular forces between the solvent and
monomer subunits dominate over intramolecular interactions. In a bad solvent or poor
solvent, intramolecular forces dominate and the chain contracts. In the theta solvent, or
the state of the polymer solution where the value of the second virial coefficient becomes
0, the intermolecular polymer-solvent repulsion balances exactly the intramolecular
monomer-monomer attraction. Under the theta condition (also called the Flory
condition), the polymer behaves like an ideal random coil.

[edit] Inclusion of plasticizers

Inclusion of plasticizers tends to lower Tg and increase polymer flexibility. Plasticizers


are generally small molecules that are chemically similar to the polymer and create gaps
between polymer chains for greater mobility and reduced interchain interactions. A good
example of the action of plasticizers is related to polyvinylchlorides or PVCs. A uPVC,
or unplasticized polyvinylchloride, is used for things such as pipes. A pipe has no
plasticizers in it, because it needs to remain strong and heat-resistant. Plasticized PVC is
used for clothing for a flexible quality. Plasticizers are also put in some types of cling
film to make the polymer more flexible.

[edit] Chemical properties

The attractive forces between polymer chains play a large part in determining a polymer's
properties. Because polymer chains are so long, these interchain forces are amplified far
beyond the attractions between conventional molecules. Different side groups on the
polymer can lend the polymer to ionic bonding or hydrogen bonding between its own
chains. These stronger forces typically result in higher tensile strength and higher
crystaline melting points.

The intermolecular forces in polymers can be affected by dipoles in the monomer units.
Polymers containing amide or carbonyl groups can form hydrogen bonds between
adjacent chains; the partially positively charged hydrogen atoms in N-H groups of one
chain are strongly attracted to the partially negatively charged oxygen atoms in C=O
groups on another. These strong hydrogen bonds, for example, result in the high tensile
strength and melting point of polymers containing urethane or urea linkages. Polyesters
have dipole-dipole bonding between the oxygen atoms in C=O groups and the hydrogen
atoms in H-C groups. Dipole bonding is not as strong as hydrogen bonding, so a
polyester's melting point and strength are lower than Kevlar's (Twaron), but polyesters
have greater flexibility.

Ethene, however, has no permanent dipole. The attractive forces between polyethylene
chains arise from weak van der Waals forces. Molecules can be thought of as being
surrounded by a cloud of negative electrons. As two polymer chains approach, their
electron clouds repel one another. This has the effect of lowering the electron density on
one side of a polymer chain, creating a slight positive dipole on this side. This charge is
enough to attract the second polymer chain. Van der Waals forces are quite weak,
however, so polyethene can have a lower melting temperature compared to other
polymers.

[edit] Standardized polymer nomenclature


There are multiple conventions for naming polymer substances. Many commonly used
polymers, such as those found in consumer products, are referred to by a common or
trivial name. The trivial name is assigned based on historical precedent or popular usage
rather than a standardized naming convention. Both the American Chemical Society[28]
and IUPAC[29] have proposed standardized naming conventions; the ACS and IUPAC
conventions are similar but not identical.[30] Examples of the differences between the
various naming conventions are given in the table below:

Common Name ACS Name IUPAC Name

Poly (ethylene oxide)


poly(oxyethylene) poly(oxyethene)
or (PEO)

Poly (ethylene poly


poly (oxy-1,2-ethanediyloxycarbonyl
terephthalate) or (oxyetheneoxyterephth=
-1,4-phenylenecarbonyl)
(PET) aloyl)

poly[amino(1-oxohexan-
Nylon poly[amino(1-oxo-1,6-hexanediyl)]
1,6-diyl)]

In both standardized conventions, the polymers' names are intended to reflect the
monomer(s) from which they are synthesized rather than the precise nature of the
repeating subunit. For example, the polymer synthesized from the simple alkene ethene is
called polyethylene, retaining the -ene suffix even though the double bond is removed
during the polymerization process:

[edit] Polymer characterization


The characterization of a polymer requires several parameters which need to be specified.
This is because a polymer actually consists of a statistical distribution of chains of
varying lengths, and each chain consists of monomer residues which affect its properties.

A variety of lab techniques are used to determine the properties of polymers. Techniques
such as wide angle X-ray scattering, small angle X-ray scattering, and small angle
neutron scattering are used to determine the crystalline structure of polymers. Gel
permeation chromatography is used to determine the number average molecular weight,
weight average molecular weight, and polydispersity. FTIR, Raman and NMR can be
used to determine composition. Thermal properties such as the glass transition
temperature and melting point can be determined by differential scanning calorimetry and
dynamic mechanical analysis. Pyrolysis followed by analysis of the fragments is one
more technique for determining the possible structure of the polymer. Thermogravimetry
is a useful technique to evaluate the thermal stability of the polymer. Detailed analyses of
TG curves also allow us to know a bit of the phase segregation in polymers. Rheological
properties are also commonly used to help determine molecular architecture (molecular
weight, molecular weight distribution and branching)as well as to understand how the
polymer will process, through measurements of the polymer in the melt phase. Another
Polymer characterization technique is Automatic Continuous Online Monitoring of
Polymerization Reactions (ACOMP) which provides real-time characterization of
polymerization reactions. It can be used as an analytical method in R&D, as a tool for
reaction optimization at the bench and pilot plant level and, eventually, for feedback
control of full-scale reactors. ACOMP measures in a model-independent fashion the
evolution of average molar mass and intrinsic viscosity, monomer conversion kinetics
and, in the case of copolymers, also the average composition drift and distribution. It is
applicable in the areas of free radical and controlled radical homo- and copolymerization,
polyelectrolyte synthesis,heterogeneous phase reactions, including emulsion
polymerization, adaptation to batch and continuous reactors, and modifications of
polymers.[31][32][33]

[edit] Polymer degradation

A plastic item with thirty years of exposure to heat and cold, brake fluid, and sunlight.
Notice the discoloration, swollen dimensions, and tiny splits running through the material

Polymer degradation is a change in the properties—tensile strength, colour, shape, etc.—


of a polymer or polymer-based product under the influence of one or more environmental
factors, such as heat, light, chemicals and, in some cases, galvanic action. It is often due
to the hydrolysis of the bonds connecting the polymer chain, which in turn leads to a
decrease in the molecular mass of the polymer. These changes may be undesirable, such
as changes during use, or desirable, as in biodegradation or deliberately lowering the
molecular mass of a polymer. Such changes occur primarily because of the effect of these
factors on the chemical composition of the polymer. Ozone cracking and UV degradation
are specific failure modes for certain polymers.
A recent finding is that polymer degradation may occur through galvanic action. In 1990,
Michael Faudree discovered that imide-linked resins in CFRP (carbon fiber reinforced
polymers) composites degrade when bare composite is coupled with an active metal in
saline, i.e. salt water environments.[34] Polymers affected include bismaleimides (BMI),
condensation polyimides, triazines, and blends thereof. Degradation occurs in the form of
dissolved resin and loose fibers. Hydroxyl ions are generated at the graphite cathode
attacking the O-C-N bond in the polyimide structure. This phenomenon, that polymers
can undergo galvanic corrosion like metals do has been referred to as the “Faudree
Effect”. Standard corrosion protection procedures were found to prevent polymer
degradation under most conditions.

The degradation of polymers to form smaller molecules may proceed by random scission
or specific scission. The degradation of polyethylene occurs by random scission—a
random breakage of the linkages (bonds) that hold the atoms of the polymer together.
When heated above 450 °C it degrades to form a mixture of hydrocarbons. Other
polymers—like polyalphamethylstyrene—undergo specific chain scission with breakage
occurring only at the ends. They literally unzip or depolymerize to become the
constituent monomer.

However, the degradation process can be useful from the viewpoints of understanding the
structure of a polymer or recycling/reusing the polymer waste to prevent or reduce
environmental pollution. Polylactic acid and polyglycolic acid, for example, are two
polymers that are useful for their ability to degrade under aqueous conditions. A
copolymer of these polymers is used for biomedical applications, such as hydrolysable
stitches that degrade over time after they are applied to a wound. These materials can also
be used for plastics that will degrade over time after they are used and will therefore not
remain as litter.

[edit] Product failure

Chlorine attack of acetal resin plumbing joint

In a finished product, such a change is to be prevented or delayed. Failure of safety-


critical polymer components can cause serious accidents, such as fire in the case of
cracked and degraded polymer fuel lines. Chlorine-induced cracking of acetal resin
plumbing joints and polybutylene pipes has caused many serious floods in domestic
properties, especially in the USA in the 1990s. Traces of chlorine in the water supply
attacked vulnerable polymers in the plastic plumbing, a problem which occurs faster if
any of the parts have been poorly extruded or injection moulded. Attack of the acetal
joint occurred because of faulty moulding leading to cracking along the threads of the
fitting, which are serious stress concentrations.

Ozone cracking in natural rubber tubing

Polymer oxidation has caused accidents involving medical devices. One of the oldest
known failure modes is ozone cracking caused by chain scission when ozone gas attacks
susceptible elastomers such as natural rubber and nitrile rubber. They possess double
bonds in their repeat units which are cleaved during ozonolysis. Cracks in fuel lines can
penetrate the bore of the tube and cause fuel leakage. If cracking occurs in the engine
compartment, electric sparks can ignite the gasoline and can cause a serious fire.

Fuel lines can also be attacked by another form of degradation: hydrolysis. Nylon 6,6 is
susceptible to acid hydrolysis, and in one accident, a fractured fuel line led to a spillage
of diesel into the road. If diesel fuel leaks onto the road, accidents to following cars can
be caused by the slippery nature of the deposit, which is like black ice.

You might also like